cond-mat0210379/qfk.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %
3: %Dear Editor, 
4: % 
5: %please find enclosed the manuscript "Quantum phase transition 
6: %in the Frenkel-Kontorova chain: from pinned instanton glass to 
7: %sliding phonon gas" which we submit for publication in Phys. Rev. E
8: %as a regular article. 
9: %The paper contains 18 figures attached in ps format. 
10: %
11: %All correspondence should be addressed to:
12: %Dima Shepelyansky
13: %Laboratoire de Physique Quantique
14: %UMR 5626 du CNRS
15: %Universite Paul Sabatier
16: %31062 Toulouse Cedex 4 
17: %France
18: %
19: %e-mail: dima@irsamc.ups-tlse.fr
20: %fax: +33-5-61556065
21: %tel: +33-5-61556068
22: %http://w3-phystheo.ups-tlse.fr/~dima
23: %
24: %Sincerely yours, 
25: %
26: %Oleg Zhirov, Giulio Casati and  Dima Shepelyansky 
27: %
28: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
29: \documentclass[twocolumn,english,aps,pre,showpacs]{revtex4}
30: \usepackage[T1]{fontenc}
31: \usepackage[latin1]{inputenc}
32: \usepackage{float}
33: \usepackage{amsmath}
34: \usepackage{graphicx}
35: \usepackage{amssymb}
36: %\usepackage[active]{srcltx} 
37: 
38: \makeatletter
39: 
40: \newcommand{\Ci}[1]{ }
41: \newcommand{\Cf}[1]{ }
42: %\newcommand{\Ci}[1]{\textbf{C$_i$(#1)}}
43: %\newcommand{\Cf}[1]{\textbf{C$_f$(#1)}}
44: 
45: \usepackage{babel}
46: \makeatother
47: \begin{document}
48: 
49: \preprint{preprint}
50: 
51: 
52: \title{Quantum phase transition in the Frenkel-Kontorova chain:\\
53: from pinned instanton glass to sliding phonon gas}
54: 
55: 
56: \author{O.V. Zhirov}
57: \email{zhirov@inp.nsk.su}
58: \affiliation{Budker Institute of Nuclear Physics, 630090 Novosibirsk, Russia}
59: 
60: \author{G. Casati}
61: \email{giulio.casati@uninsubria.it}
62: \affiliation{International Center for the Study of Dynamical 
63: Systems, Universit\`a degli Studi dell'Insubria and\\
64: Istituto Nazionale per la Fisica della Materia, 
65: Unit\`a di Como, Via Valleggio 11, 22100 Como, Italy and\\
66: Istituto Nazionale di Fisica Nucleare, 
67: Sezione di Milano, Via Celoria 16, 20133 Milano, Italy}
68: 
69: \author{D.L. Shepelyansky}
70: \email{dima@irsamc.ups-tlse.fr}
71: \affiliation{Laboratoire de Physique Quantique, UMR 5626 du CNRS, Universit\'{e}
72: Paul Sabatier, 31062 Toulouse, France}
73: 
74: 
75: \date{October 17, 2002}
76: 
77: \begin{abstract}
78: We study analytically and numerically the one-dimensional quantum Frenkel-Kontorova
79: chain in the regime when the classical model is located in the pinned 
80: phase characterized by the gaped phonon excitations and devil's staircase.
81: By extensive quantum Monte Carlo simulations we show 
82: that for the effective  Planck constant $\hbar$  
83: smaller than the critical value
84: $\hbar_c$ the quantum chain is in the pinned instanton glass phase.
85: In this phase
86: the elementary excitations have two branches:
87: \emph{phonons}, separated from zero energy by a finite gap, and \emph{instantons} which
88: have an exponentially small excitation energy. At  $\hbar=\hbar_c$
89: the quantum phase transition
90: takes place and for $\hbar>\hbar_c$ the pinned instanton glass is transformed into the
91: sliding phonon gas with gapless phonon excitations.
92: This transition is accompanied by the divergence of the spatial correlation
93: length and appearence of sliding modes at $\hbar>\hbar_c$.
94: \end{abstract}
95: \pacs{05.45.-a, 63.70.+h, 61.44.Fw }
96: \maketitle
97: 
98: 
99: 
100: \section{introduction}
101: 
102: The Frenkel-Kontorova (FK) model \cite{FK38} describes a one-dimensional
103: chain of atoms/particles with harmonic couplings placed in a periodic
104: potential. This model was introduced more than sixty years ago with
105: the aim to study crystal dislocations \cite{FK38,Naba67}. It was
106: also successfully applied later to the description of commensurate-incommensurate
107: phase transitions \cite{VPaAT84}, epitaxial monolayers on the crystal
108: surface \cite{SY71}, ionic conductors and glassy materials \cite{Piet81,Au78,Au83a}
109: and, more recently, to charge-density waves \cite{Flor96} and dry
110: friction \cite{Brau97,Cons00}. Despite the fact that the relevant
111: phenomena are at the atomic scale, all these works are based essentially
112: on the classical approach. The first study of quantum effects was done 
113: twelve years ago \cite{Borg89,Borg90,Borg89d} with the attempt
114: to understand the highly nontrivial quantum ground state of the Frenkel-Kontorova model
115: in the regime when the classical ground state is characterized
116: by the fractal ``devil staircase'' \cite{Au78}. These studies were
117: extended in 
118: \cite{Berm94,Berm97} and later the quantum dynamics at different values
119: of quantum parameter $\hbar $ was studied in \cite{Hu98,Hu99,Hu01,Hu01r,Ho01}. 
120: 
121: 
122: The physical properties even of the classical FK model are very rich 
123: and nontrivial. In 1978 Aubry discovered \cite{Au78} a new type of ground
124: state which has fractal properties known as ``devil's staircase''. 
125: In fact, the equilibrium positions of atoms in the FK chain 
126: are described \cite{Au78} by the well known Chirikov standard map \cite{Chir79},
127: which describes generic properties of chaotic Hamiltonian dynamics.
128: The density of particles in the equilibrium
129: state determines the rotation number of the invariant curves of the
130: map, while the amplitude of the periodic potential in the FK model
131: gives the value of the dimensionless parameter $K$. According to
132: the known properties of the map, it follows that there exist two phases
133: of the chain, the ``sliding'' phase and the ``pinned''
134: phase. Indeed at $K<K_{c}$ the Kolmogorov-Arnold-Moser (KAM) curves are smooth and 
135: the chain can easily slide along the potential. This implies the existence of
136: a zero phonon mode. In contrast, at $K>K_{c}$ the KAM curves are
137: destroyed and replaced by an invariant Cantor set which is called
138: cantorus \cite{Au78,Perc79,Au83b,Au83c,Mac84}. In this ``pinned'' phase the chain cannot
139: slide being kept by a finite Peierls-Nabarro barrier and the phonon
140: spectrum is separated from zero by a finite gap. In this paper we consider only
141: the gaped/pinned phase with  $K>K_{c}$.
142: 
143: It is important to stress that in the pinned phase, besides the equilibrium state
144: of minimal energy, there exists a lot of other \emph{equilibrium} configurations,
145: corresponding to \emph{local} minima of the potential with energies
146: very close to the minimal energy, i.e. the energy of the ground state
147: \cite{Sch84,Sch85,Zhir01a}. The total amount of these states (configurational
148: excitations) grows exponentially both with the length $L$ of the
149: chain, as well as with the parameter $K$ \cite{Zhir01a}. Moreover, a great number of
150: them is practically degenerate since their energy separation from
151: the ground state is exponentially small. 
152: 
153: In the classical limit all these configurational excitations
154: are stable, while in the quantum case they become metastable due to tunneling
155: between different exponentially degenerate minima of the potential. As 
156: a result one may expect that the true eigenstates of the Hamiltonian, including
157: the ground state, are built as superpositions of classical configurational
158: states. Such a nontrivial structure of the quantum ground state of the chain has
159: a profound  analogy with the famous vacuum of Quantum
160: Chromodynamics (QCD) \cite{Shur98} in which tunneling transitions between different,
161: practically degenerate, states are known as ``instantons''.
162: This analogy between the two problems is very useful and implies 
163: that many of the methods developed in lattice QCD studies may
164: be applicable to quantum FK model.
165: 
166: Dynamical low energy excitations in the classical FK chain consist
167: only of phonon modes describing small vibrations around a classical minimum
168: of the chain potential energy. In the quasiclassical regime, where  
169: the effective dimensionless Planck constant $\hbar$ is very small,
170: the quantization of the FK chain can be reduced to a quantization of phonon
171: modes only. Indeed the time of tunneling between different minima of
172: the potential energy is exponentially large, compared to periods of
173: the vibrations, and the phonon modes are decoupled from the tunneling
174: modes (instantons). 
175: This reminds the situation in the QCD \cite{Shur98}
176: where the quasiclassical regime for instantons appears at small distances
177: on which instantons are also decoupled from other excitations like 
178: quarks and gluons.
179: 
180: In this regime the tunneling transitions are very slow, instantons are local
181: and frozen in space. Due to exponential degeneracy of chain configurations the
182: model reveals a glass-like structure of instantons randomly distributed   
183: along the chain. We will call this phase the ``instanton glass''.
184: 
185: One may expect that the phase structure changes significantly with the increase 
186: of $\hbar$ when the
187: tunneling time $t\sim \exp (S_{in}/\hbar )$ ($S_{in}$ is the instanton
188: action) becomes comparable to the inverse frequency of phonons. In this
189: case phonon and instanton excitations can strongly influence each other
190: due to anharmonicity of the periodical
191: potential. This may lead to quantum melting of the instanton glass phase
192: and transition to another phase which appears above some critical value 
193: $\hbar>\hbar_c$.
194: 
195: The existence of another quantum phase can be argued in the following way.  
196: At sufficiently large  $\hbar \geq \hbar _{c}$ the kinetic energy of
197: quantum particle $E_{k}\sim \hbar^{2}/2m(\Delta x)^{2}$ starts to exceed
198: the height $U$ of the periodic potential in which the chain is placed
199: (here $\Delta x$ is the mean particle separation in the chain which is comparable
200: with the period of the potential, $m$ is the particle mass). 
201: The condition $E_{k}\sim U$ gives $\hbar_c\sim\Delta x\sqrt{mU}$.
202: As a result, for $\hbar>\hbar_c$ the chain turns from the \emph{pinned} 
203: to the \emph{sliding} phase. In this regime tunneling is replaced by direct 
204: propagation above the barrier.
205: Hence, the instanton-like motion is replaced by phonon-like motion corresponding 
206: to a new phase with gapless phonon excitations. This regime can be considered as 
207: sliding quantum phase similar to classical sliding regime at $K<K_c$ \cite{Au83b,Au83c}.
208: Qualitative different types of the behavior of quantum FK model at small and large
209: $\hbar$ values were already seen in the first numerical studies \cite{Borg89,Borg90}.  
210: In summary, at $\hbar =\hbar _{c}$
211: one may expect a \emph{quantum phase transition}, with qualitative
212: rearrangement of the spectrum of elementary quasiparticle excitations.
213: 
214: 
215: In this work, devoted to a comprehensive numerical study of the transition
216: we have just outlined, we present a complete picture of low energy
217: excitations in the quantum FK chain. We also find that this
218: transition is highly nontrivial and the model reveals a non-analytical
219: behavior which can partially explain the failure of simple analytical
220: approaches developed in \cite{Berm94,Hu98,Hu99,Ho01}. It should be
221: stressed also that our results have relevance to a wider class of
222: quantum systems, like e.g. quantum spin glass or other disordered systems 
223: with interactions. Indeed, the existence of highly degenerate classically 
224: stable configurations is a general property of such systems.
225: 
226: 
227: The paper is organized as follows. In Section \ref{sec:the-fk-model}
228: we outline the model, its quantum features and the main points of our
229: numerical approach. In Section \ref{sec:elementary-excitations.}
230: we study  elementary excitations of the chain at different
231: values of $\hbar $ and show that there exists a structural rearrangement of
232: the excitation spectrum at certain $\hbar = \hbar_c$. 
233: A detailed analysis of this rearrangement, given in Section \ref{sec:QFT}, 
234: indicates that we have a quantum phase transition. Our results are 
235: summarized in Section \ref{sec:Conclusions}.
236: 
237: 
238: \section{The quantum Frenkel-Kontorova model}
239: \label{sec:the-fk-model}
240: 
241: \subsection{Definitions and outline of quantum features}
242: 
243: The model describes a one-dimensional chain of particles with harmonic
244: couplings placed in a periodic potential. The Hamiltonian reads:
245: \begin{equation}
246:   H=\sum _{i=0}^{s}\left[\frac{P_{i}^{2}}{2m}+\alpha\frac{(x_{i}-x_{i-1})^{2}}{2}-
247:   \beta\cos(x_{i}/d)\right]
248: \label{H_FKorig}
249: \end{equation}
250:  where $s$ is the number of particles, $P_{i}$ and $x_{i}$ are
251: their momenta and coordinates, and in the quantum case 
252: $P_{i}=-i\hbar\partial/\partial x_i$. In this paper we use  
253: units $m=d=\alpha=1$, that correspond to dimensionless Hamiltonian 
254: (see for details e.g. \cite{Borg89,Borg90,Berm94}):
255: \begin{equation}
256:   H=\sum _{i=0}^{s}\left[\frac{P_{i}^{2}}{2}+\frac{(x_{i}-x_{i-1})^{2}}{2}-
257:   K\cos(x_{i})\right]
258: \label{H_FK}
259: \end{equation}
260: with a dimensionless parameter $K=\beta/\alpha d^2$ and the
261: dimensionless Planck constant $\hbar$ is measured in units of $d^2\sqrt{m \alpha}$. 
262: In this way, $K$ is the chaos parameter in the Chirikov standard map \cite{Chir79,Au83b}.
263: We use the standard boundary conditions
264: \begin{equation}
265:   x_{0}=0,\quad x_{s}=L\, ,\label{Bcond}
266: \end{equation}
267: where the chain length $L=2\pi \cdot r$ consists of $r$ periods/wells 
268: of the external field. We analyze the standard case of golden mean ratio corresponding 
269: to $r/s\to(\sqrt{5}-1)/2$ (see, \cite{Borg89,Borg90,Zhir01a}).
270: The potential energy of the chain
271: \begin{equation}
272:   U(\{x\})=\sum _{i=0}^{s}\left[\frac{(x_{i}-x_{i-1})^{2}}{2}-
273:   K\cos (x_{i})\right]
274: \label{Upot}
275: \end{equation}
276: has a large number of minima, corresponding to different possible
277: distributions of particles among the wells. The classical 
278: ground state (absolute minimum of the potential energy) is
279: characterized by some special ordering discovered by Aubry 
280: \cite{Au78,Au83b,Au83c}. In addition, there are also local
281: minima, known as ``configurational excitation'' states\cite{Be80}.
282: Many of them have exponentially small energy separations from
283: the energy of the classical ground state \cite{Zhir01a}. In the classical
284: case all these states are well defined (distinguishable) and absolutely
285: stable.
286: 
287: In contrast, in the \emph{quantum} world these states are \emph{metastable}
288: due to quantum tunneling. Since they are practically degenerate with
289: the classical ground state the actual \emph{quantum} ground state
290: is built by \emph{many} of them. The admixture of metastable classical 
291: configurations in the quantum ground state was first discussed in \cite{Berm94,Berm97}. 
292: Let us summarize below 
293: the most important aspects of the quantum ground state, related to
294: the tunneling between these configurations.
295: 
296: In the quasiclassical region, the transition amplitudes between different
297: metastable configurations are exponentially small. Hence,
298: the ground state wave function is (with exponential accuracy) a sum
299: of \emph{non}-overlapping parts, each referring to a particular classical
300: configuration. Any average over the quantum ground state is (within
301: the same accuracy) a weighted sum of averages over relevant classical
302: configurations. Note that in this limit the main contribution to quantum
303: motion comes from phonons which characterize small vibrations around a 
304: classical equilibrium 
305: configuration. In fact, the phonon spectrum is only weakly sensitive
306: to the choice of specific configuration \cite{PhonFK} and 
307: the influence of tunneling processes on the global (thermodynamical) 
308: properties of the chain is expected to be small. In particular,
309: the phonon spectrum is still characterized by a phonon gap similar to 
310: the classical case. In this regime the tunneling 
311: transitions between metastable configurations are very slow comparing to 
312: phonon frequencies and can be considered as well separated instantons.
313: 
314: At higher values of $\hbar $ the tunneling rate increases and becomes
315: comparable with the frequency of phonons. In this regime phonon
316: oscillations are large and essentially anharmonic and therefore the
317: known analytical approaches \cite{Berm94,Hu98,Hu01r,Ho01} based on
318: the gauss-like wave function profile are not quite adequate. Instead, 
319: interactions between instantons and phonons come into play here.
320: As it will be shown later,  this leads to a new sliding phase appearing
321: at $\hbar>\hbar_c$. 
322: 
323: In fact, the first effects of quantum tunneling between classical configurations
324: were already seen in the numerical studies \cite{Borg90} (see Fig.4b). However, 
325: they were not attentively analyzed and the mixture of classical metastable 
326: configurations induced by quantum tunneling was not discussed. 
327: Certainly, tunneling processes affect significantly the average positions of 
328: particles at large time scales. Therefore the quantities based on mean expectation 
329: values of particle positions are not adequate even in the deep quasiclassical regime. 
330: Hence the basic concepts of classical treatment, like hull \cite{Au78,Au83b}
331: and g-functions \cite{Borg89,Borg90}, are also not adequate in this quantum regime.
332: The point is that in the case of particle tunneling between
333: two classical equilibrium positions the expectation value
334: does not correspond to any \emph{probable} (in the classical sense)
335: particle position. This is evident from the analogy with the two-well
336: potential problem, where the mean expectation value of a particle
337: position coincides with the \emph{top} of the barrier. 
338: 
339: 
340: \subsection{Path integral and numerical simulations}
341: \label{sub:Path-Sim}
342: 
343: Tunneling effects are best understood in the Feynman path integral
344: formulation of quantum mechanics\cite{Poly77}. In particular, the
345: transformation to ``Euclidean'' time variable $\tau$: $t=i\tau$
346: makes the tunneling transition, or \emph{instanton,} local in time
347: $\tau$. In the quasiclassical regime the tunneling probability
348: is very small and therefore the mean separation between instantons 
349: is large compared to their size. This allows to use the approximation
350: of dilute instanton gas \cite{Poly77}. 
351: 
352: Feynman path formulation in the ``Eucleudean'' time 
353: allows direct numerical simulations of quantum systems with
354: probabilistic treatment \cite{Creu81} of the path integral:
355: \begin{equation}
356:   Z=\int Dx[\tau ]\exp (-\frac{1}{\hbar }S[x(\tau )]),
357: \label{FeyP}
358: \end{equation}
359: where the ``Eucledian'' action
360: \begin{equation}
361:   S[x(\tau )]=\int _{0}^{\tau _{0}}d\tau \sum_i
362:   \left(\frac{\dot{x}_{i}^{2}}{2}+\frac{(x_{i}-x_{i-1})^{2}}{2}-
363:   K\cos (x_{i})\right)
364: \label{Sc}
365: \end{equation}
366: has the same form as a total energy of a chain of particles in the usual 
367: real time variable, integrated over some time interval $\left[0,\tau _{0}\right]$.
368: This is equivalent to the consideration of the quantum system at some finite
369: temperature\cite{Creu81}
370: \begin{equation}
371:   T=\hbar /\tau _{0}  %\label%{eq:Tempr}
372: \label{eq:Temperature} 
373: \end{equation}
374: 
375: We assume periodic boundary conditions in the \emph{time}
376: direction, which correspond to a path closed on a torus 
377: \begin{equation}
378:   x_{i}(\tau _{0})=x_{i}(0),\quad i=1,\ldots ,s-1,
379: \label{TbndCnd}
380: \end{equation}
381: and integrate over the initial conditions $x_{i}(0)$ in order to
382: restore homogeneity of paths along the time torus. This allows to
383: improve the statistics in data measurements by averaging the data
384: along the torus.
385: 
386: The numerical simulation of the path integral (\ref{FeyP}) needs
387: discretization of the time variable $\tau _{n}=\Delta \tau \cdot n$,
388: $n=1,\ldots ,N$ by splitting the time interval $\tau_ {0}$ into $N$
389: steps of size $\Delta \tau =\tau_0 /N$. As a result the original 
390: time-continuous model (\ref{H_FK}) turns into a 2-dimensional lattice 
391: model with the action
392: \begin{eqnarray}
393:   S & = & \sum _{n=0}^{N}\sum _{i=1}^{s}\left[\right.
394:         \frac{(x_{i,n+1}-x_{i,n})^{2}}{2\Delta \tau }+\nonumber \\
395:     &   & +\Delta \tau \frac{(x_{i+1,n}-x_{i,n})^{2}}{2}-
396:         \Delta \tau \cdot K\cos x_{i,n}\left.\right].
397: \label{eq:S_Lat}
398: \end{eqnarray}
399: 
400: The time step $\Delta \tau $ should be chosen small enough to approach
401: the continuous limit of the original model (\ref{H_FK}). This leads
402: to the following requirements. At any time step $\Delta \tau $
403: the path variable $x_{i}$ jumps by a random shift 
404: $\Delta x_{i}\sim \sqrt{\hbar \Delta \tau }$. The first obvious requirement 
405: is that the shift $\Delta x_{i}$ should be small compared to the spatial scale 
406: of the potential energy variation.
407: Another requirement comes from the standard derivation of the path
408: integral \cite{Creu81,Shur84}: the potential energy terms should
409: be small compared to the kinetic energy:
410: \begin{equation}
411:   \Delta \tau \cdot \left(\frac{(x_{i+1,n}-x_{i,n})^{2}}{2}-K\cos x_{i,n}\right)
412:   \ll \frac{(x_{i,n+1}-x_{i,n})^{2}}{2\Delta \tau }.
413: \label{P<<K}
414: \end{equation}
415: The latter condition makes our lattice \emph{anisotropic} with respect
416: to spatial/time directions. In spite of the common belief \cite{cQFT97}
417: on the equivalence between quantum 1-dimensional chains and classical
418: 2-dimensional statistical (e.g. \textsl{XY}) models, the above discussion
419: shows that some particular care has to be taken.
420: 
421: An accurate treatment of particle tunneling between two wells in the chain
422: requires some modifications of the action (\ref{eq:S_Lat}).
423: For a finite (not very small!) time step $\Delta \tau $, there exists
424: a probability for a particle to jump over the potential barrier in
425: \emph{one} time step\cite{Shur84}. The corresponding path contribution
426: to the action will be strongly underestimated by eq.(\ref{eq:S_Lat})
427: and, respectively, the probability of tunneling will be too high.
428: To cure this problem we use in our simulation an improved \cite{Shur84}
429: version of action (\ref{eq:S_Lat}), with the potential energy term
430: $\Delta \tau \cdot U(x_{n})$ replaced by the integral 
431: $\Delta \tau \cdot \int _{x_{n}}^{x_{n+1}}U(x)dx/(x_{n+1}-x_{n})$
432: along a straight line that links subsequent (in time) points $(x_{n},x_{n+1})$
433: of the particle path. This significantly improves the accuracy of 
434: numerical simulations.
435: 
436: For simulations of path ensembles we use the standard Metropolis algorithm
437: \cite{Metr53}. Each iteration looks as follows: at any fixed time
438: slice at number $n$ we update sequentially the particles coordinates
439: $x_{i,n}$, $i=1,\ldots ,s$; then we go to the next time slice: $n\rightarrow n+1,$
440: and so on. 
441: 
442: The system has two principal characteristic relaxation time scales
443: that are originated by different underlying processes. Basically one
444: can estimate the number of iterations as $N_{it}\sim 1/\omega ^{(min)}\Delta \tau $,
445: where $\omega^{(min)}$ is the lowest frequency relevant
446: to the process and $\Delta \tau $ is time discretization step. The
447: shortest scale is related to the path relaxation with respect
448: to main phonon modes, and is typically of the order of a few tens of iterations
449: in the quasiclassical regime, where phonons have a gap of
450: order unity. Another time scale is much larger and is determined by the smallest
451: frequency related to either the tunneling rate between
452: different classical configurations, or to the lowest phonon
453: frequency available in the system for $\hbar \geq \hbar _{c}$.
454: 
455: For the initial state we choose the Aubry classical ground state. Then, applying 
456: iterations, we generate a path ensemble. To be sure that the system does actually 
457: relax to statistical equilibrium with respect to the slowest processes described above, 
458: we control the mean number of particle path crosses a top of the potential 
459: barrier and we discard all configurations in the ensemble until this quantity
460: stabilizes. All computations are done for the chaos parameter $K=5$.
461: The required number of iterations to reach relaxation is very sensitive 
462: to the value of quantum parameter $\hbar $: for example at $\hbar =3$ this number is
463: $N_{it}\sim (2\div 4)10^{2}$ , while  at $\hbar=1$ it is  of the order of $10^{5\div 6}$.
464: This explains why in the first studies \cite{Borg89,Borg90} done at
465: $N_{it}\leq10^4$ many details at $\hbar \leq 2$ were not seen.
466: We study chains with up to 233 particles.
467: 
468: \section{elementary excitations}
469: \label{sec:elementary-excitations.}
470: 
471: The most important information about the quantum system is contained
472: in its spectrum of low-lying elementary excitations. Being the net
473: manifestation of system internal structure it reflects any structural
474: transition which can occur in the system, and it provides a complete
475: description of low-temperature thermodynamic and kinetic
476: properties. We extract this spectrum using a novel approach based
477: on the analysis of Fourier spectrum of Feynman paths. This approach
478: provides a most direct way to see and resolve different excitations
479: in the system. Then  we compare our method with the more traditional
480: one, based on the study of time correlation functions. This comparison 
481: provides a self-consistency check and demonstrates the advantages 
482: of our method.
483: 
484: 
485: \subsection{Spectral properties of Feynman paths}
486: \label{sub:SpFunc}
487: 
488: In classical nonlinear dynamics, the Fourier analysis of trajectories plays
489: a key role in understanding of periodic motion of complex systems. In
490: a similar way,
491: the spectral characteristics of Feynman paths are closely related
492: to the properties of elementary excitations in quantum systems. We start
493: our studies from the quasiclassical limit $\hbar \to 0$, where
494: this relation is exact, and extend them to higher values of
495: $\hbar $.
496: 
497: Let us consider the Fourier image of the path variable $x_{i}(\tau )$:
498: \begin{equation}
499:   a_{i}(\omega _{m}) = \frac{1}{\sqrt{\tau _{0}}}\int _{0}^{\tau _{0}}d\tau 
500:   \, x_{i}(\tau )\, \exp (i\omega _{m}\tau )
501: \label{eq:four_x}
502: \end{equation}
503: where $\omega _{m}=m\varpi $, $\varpi \equiv 2\pi /\tau _{0}$, and
504: $-\infty <m<\infty $. The path variable $x_{i}(\tau )$ is real,
505: therefore $a_{i}(-\omega _{m})=(a_{i}(\omega _{m}))^{\ast }$. To
506: get insight into the physical content of this quantity let us consider
507: the quasiclassical regime $\hbar \ll 1$. Then, for small variations
508: $x_{i}(\tau )=\overline{x}_{i}+\delta x_{i}(\tau )$ around the classical
509: static trajectory $\{\overline{x}_{i}\}$, one can expand the action
510: (\ref{Sc}) up to second order terms in $\delta x_{i}(\tau )$. Next,
511: using the spectral expansion for $\delta x_{i}(\tau )$ and performing
512: the integration over $\tau $, one gets:
513: \begin{eqnarray}
514:  S & = & S_{0}[\overline{x}]+\nonumber \\
515:    & + & \sum _{m}\sum _{i,k}\frac{1}{2}(\omega _{m}^{2}\delta _{ik}-
516:    \Omega _{ik}^{2})a_{i}(-\omega _{m})a_{k}(\omega _{m})\label{eq:Sexpa}\\
517:    \Omega _{ik}^{2} & = & \delta _{ik}\left(2+K\cos (x_{i})\right)-
518:    \delta _{i,k-1}-\delta _{i-1,k}\label{eq:Om2}
519: \end{eqnarray}
520: Now, by the transformation to normal modes 
521: $A^{(l)}(\omega _{m})=\sum _{i}V_{i}^{(l)}a_{i}(\omega _{m})$,
522: where eigenvectors $V_{i}^{(l)}$ satisfy the equation 
523: $\Omega _{ik}^{2}V_{k}^{(l)}=\nu _{l}^{2}V_{i}^{(l)}$
524: one gets the standard representation of the action as a sum of independent
525: phonon modes:
526: \begin{equation}\label{eq:Om3}
527: S=S_{0}[\overline{x}]+\sum _{m}\sum _{l}\frac{1}{2}(\omega _{m}^{2}+\nu _{l}^{2})\left|A^{(l)}(\omega _{m})\right|^{2}
528: \end{equation}
529: Finally, the path integral (\ref{FeyP}) turns into a product of ordinary
530: integrals
531: \begin{eqnarray*}
532: Z & = & \int \prod _{l=1}^{s-1}dA^{(l)}(0)\exp \left(-\nu _{l}^{2}\left|A^{(l)}(0)\right|^{2}/2\hbar \right)\times \\
533:  &  & \quad \prod _{m=1}^{\infty }d{\textrm{Re}}A^{(l)}(\omega _{m})d{\textrm{Im}}A^{(l)}(\omega _{m})\times \\
534:  &  & \quad \exp \left(-(\omega _{m}^{2}+\nu _{l}^{2})\left|A^{(l)}(\omega _{m})\right|^{2}/\hbar \right),
535: \end{eqnarray*}
536: and one arrives to the well known result for the correlator of free phonon
537: modes: 
538: \begin{equation}
539:  \left\langle A^{(l)}(\omega _{m})A^{(l)\ast }(\omega _{m'})\right\rangle =
540:  \frac{\hbar \delta _{mm'}}{(\omega _{m}^{2}+\nu _{l}^{2})}.
541: \label{eq:corrAA}
542: \end{equation}
543: We note that this result is obtained from Eq.(\ref{Sc}) where the
544: action is \emph{continuous} in time variable. In the discretized version
545: (\ref{eq:S_Lat}) the number of harmonics is finite: $\left|m\right|=0,1,\ldots ,M$,
546: where $M=\tau _{0}/2\Delta \tau $. It can be shown that the only modification 
547: induced by discretization is the replacement in (\ref{eq:corrAA}):
548: $\omega _{m}\rightarrow \widetilde{\omega }_{m}=(2\omega _{M}/\pi )\sin (\pi \omega _{m}/2\omega _{M})$,
549: $\omega _{M}\equiv M\varpi $. As a result, the spectral function 
550: for phonons is given by
551: \begin{equation}
552:  F^{(l)}(\widetilde{\omega }_{m})\equiv 
553:    \left\langle \left|A^{(l)}(\omega _{m})\right|^{2}\right\rangle =
554:    \frac{\hbar }{(\widetilde{\omega }_{m}^{2}+\nu _{l}^{2})}.
555: \label{eq:Fw}
556: \end{equation}
557: Hereafter, instead of $\omega_m$ we assume its discretized version 
558: $\widetilde{\omega}_m$, and in the following  the tilde will be omitted.
559: 
560: The expression (\ref{eq:corrAA}) is the well known Wick rotated 
561: Green function (in the frequency representation) for a single free particle 
562: in the phonon field theory, which has in our case one spatial dimension.
563: 
564: In the quasiclassical regime the amplitudes of phonon oscillations are small and 
565: the interactions between phonons due to anharmonicity of the Hamiltonian (\ref{H_FK}) 
566: are negligible. At higher $\hbar$, the amplitudes of phonon vibrations grow 
567: as $\hbar^{1/2}$ and their interactions become more important. In general, 
568: interactions can essentially modify the Green function (\ref{eq:corrAA}) for 
569: phonon excitations.  This actually happens for $\hbar>\hbar_c$ where the spectrum
570: of excitations is significantly changed.
571: However, one may expect that the  spectral function of elementary excitation 
572: remains of the same form 
573: \begin{equation}
574:   F^{(l)}(\omega _{m})=
575:   f\frac{\hbar }{(\omega_{m}^{2}+\nu _{l}^{2})},
576: \label{eq:Fw_renorm}
577: \end{equation}
578: which differs from (\ref{eq:corrAA}) by the renormalized frequency value $\nu _{l}$
579: and by an overall renormalization factor $f$, in analogy with the Green function 
580: behavior in renormalizable quantum field theories (see, e.g. \cite{Shur98}). 
581: In fact, this idea is well supported by numerical data.
582: 
583: An extended elementary excitation involves all particles in the chain. In turn,
584: the Fourier harmonics of any particle coordinate in the chain are a
585: sum of contributions of many elementary excitations
586: \begin{equation}
587:  \left\langle \left|a_{i}(\omega _{m})\right|^{2}\right\rangle =
588:  \sum _{j}f^{(j)}_i\frac{\hbar }{(\omega_{m}^{2}+\nu _{j}^{2})},
589: \label{eq:Fw_x_i}
590: \end{equation}
591: where the sum goes over all chain excitations.
592: In the deep quasiclassical case the main contribution comes from phonons modes, 
593: and the sum goes over phonon modes $l$, with 
594: $f^{j}_i\to\left|V_{i}^{(l)}\right|^{2}$ and $\nu_{j}\to\nu_l$.
595: 
596: The goal of our study here is a complete picture of low-lying quantum
597: excitations in the chain, both for low and high values of $\hbar $.
598: In fact, the spectrum of low-lying excitations is crucially dependent
599: on $\hbar $: there are domains with a \emph{qualitatively} different
600: behavior. In order to illustrate this, let us consider the amplitude
601: of quantum motion of particles in the chain, given by Eq.(\ref{eq:Fw_x_i}).
602: It is seen that the contributions of low-frequency modes(small $\nu_l$) dominate in the
603: limit $\omega _{m}\rightarrow 0$, provided that their wave-function
604: profiles $V_{i}^{(l)}$ are not small at the particle position $i$.
605: Therefore the spectral function (\ref{eq:Fw_x_i}), computed in this limit,
606: gives a rough estimate for the frequency $\nu_l$ of the lowest mode. 
607: 
608: %
609: \begin{figure}[th]
610: \includegraphics[  clip,
611:   width=105mm,
612:   height=85mm,
613:   keepaspectratio,
614:   angle=270,
615:   origin=c]{fig1.ps}
616: \vskip -10mm
617: \caption{\label{fig:overview}Dependence of the amplitude of the lowest Fourier
618: harmonic $a_{i}\equiv a_{i}(\omega_1)$  on the particle
619: position $i$ in the chain at different $\hbar $. 
620: Here $\omega_1=\varpi =2\pi /\tau _{0}$, $m=1$. The chain parameters
621: are $s/r=89/55$, $K=5$, $\tau _{0}=80$. Typical number of iterations 
622: is $(1.5\div 5)\cdot 10^5$ at each
623: value of $\hbar$.}
624: \end{figure}
625: 
626: 
627: In Fig.\ref{fig:overview} the amplitude of the \emph{lowest} Fourier
628: harmonic $a_{i}(\omega_1)$ with $\omega_1=\varpi =2\pi /\tau _{0}$ is plotted
629: as a function of the particle position $i$ at different $\hbar =0.6-8$.
630: One can see that the whole interval of $\hbar $ splits naturally in
631: \emph{three} regions of qualitatively different behavior: (i) the
632: quasiclassical region $\hbar \lesssim 1$ where the amplitudes $a_i(\omega_1)$ of
633: the harmonics are very small and depend on the particle positions in some regular way;
634: (ii) the transition region $1\lesssim \hbar \lesssim 2$,
635: where this dependence is highly irregular and interactions between instantons and
636: phonons are important; and (iii) the region $\hbar \gtrsim 2$,
637: where this dependence becomes regular again and where, as we shall see below, 
638: a new phonon branch appears. 
639: Let us note, that at $\hbar<1$ the regular structure along the chain
640: is quasiperiodical, which reflect a fact that a classical chain is built of ``bricks'' 
641: of two principal sizes \cite{Zhir01a}. Above $\hbar\approx 2$ bricks are ``melted''
642: and chain properties become even more homogeneous along the chain. In the 
643: intermediate region irregular peaks come from different non-overlapping
644: instantons contributions, which as 
645: any tunneling effects are highly sensitive to small variations of potential barriers.
646: Below their contributions are exponentially small, and above, as we see further, they 
647: overlap and form new \textit{sliding} phase of the system.
648: 
649: 
650: In the following we analyze in detail these regions, corresponding to different 
651: intervals of $\hbar$.
652: 
653: 
654: \subsection{Quasiclassical region $\hbar \lesssim 1$}
655: \label{sub:Quasiclassical}
656: 
657: For $\hbar\lesssim 1$ the tunneling between different 
658: metastable classical configurations is negligible, and the particles mainly 
659: vibrate around some classical equilibrium positions. In this case the elementary
660: excitations are phonons, and the quantization of the chain is reduced
661: to the quantization of phonon modes, see e.g. \cite{Hu01}. 
662: 
663: To single out low energy excitations we use the following approach.
664: Of course, a particular excitation can be selected if the corresponding mode
665: $V_{j}^{(l)}$ is known, but in general this is not a trivial task.
666: However, for low lying excitations one may expect that the modes have a simple harmonic
667: form\begin{eqnarray}
668: V_{j}(k_{l})\equiv V_{j}^{(l)} & = & \sqrt{\frac{2}{L}}\sin (k_{l}j),\label{eq:profV}\\
669:  &  & k_{l}=\pi l/L,\; (l=1,2,\ldots )\nonumber 
670: \end{eqnarray}
671: where the wavelength $\lambda _{l}\equiv 2\pi /k_{l}$ is
672: much larger than a characteristic size of inhomogeneity in the chain.
673: Direct numerical computations \cite{PhonFK} of phonon modes in the classical 
674: FK chain support this ansatz (\ref{eq:profV}).
675: 
676: %
677: \begin{figure}[ht]
678: \includegraphics[  clip,
679:   width=115mm,
680:   height=85mm,
681:   keepaspectratio,
682:   angle=90,
683:   origin=c]{fig2.ps}
684: 
685: \vskip -12mm
686: \caption{\label{fig:h08exmpla} The phonon spectral function $F^{(1)}(\omega _{m})$
687: versus the rescaled frequency $\omega _{m}$, data are shown
688: for the lowest spatial mode with $l=1$. The chain parameters are: 
689: $s/r=89/55$, $K=5$,$\tau _{0}=80$, $\hbar=0.8$. Here $m$ varies from $1$ to $89$,
690: but for clarity only selected values are shown. The solid
691: curve gives the fit by  Eqs.(\ref{eq:Fw},\ref{eq:Fw_renorm}), open circles show numerical
692: data. The fit determines the phonon frequency of the first mode ($\nu_1^2=3.170\pm0.012$)
693: and the renormalization factor $f=1.0034\pm0.0034$.}
694: \end{figure}
695: 
696: The numerical test of this anzatz (\ref{eq:profV}) is given in Fig.\ref{fig:h08exmpla}.
697: Here a typical result of quantum simulations of the spectral function
698: $F^{(l)}(\omega _{m})$ is shown for the lowest phonon
699: mode $l=1$ at $\hbar=0.8$. Fitting the data by Eq.(\ref{eq:Fw_renorm})
700: with the renormalization factor $f$ and the frequency of the
701: phonon mode $\nu _{l}$ as  free parameters, we obtain
702: $f=1.0034\pm 0.0034$ and $\nu_1^2=3.170\pm0.012$.
703: This fit shows that the renormalization factor $f$ is remarkably close
704: to unity, in spite of the fact that the value $\hbar =0.8$ is not
705: small. Hence, the ansatz (\ref{eq:profV})
706: provides a good approximation to actual profiles of lowest phonon
707: modes. This also indicates that the Gauss approximation used
708: in the Section \ref{sub:SpFunc} and in papers \cite{Hu98,Hu01r,Ho01}
709: works fine here. However we note that  at the same time the  quantum effects
710: renormalize substantially the phonon frequency ($\nu_1^2=3.170$)
711: compared to its value at $\hbar =0.2$  ($\nu _{1}^{2}=3.706\pm 0.019$).
712: The computations for the classical FK chain give $\nu _{1}^{2}=3.717$ ($\hbar=0$).
713: 
714: These data show that at small $\hbar$ the frequency $\nu_1$ obtained from
715: the quantum simulations approaches to the frequency of phonon mode in the classical chain.
716: This fact gives a very important check of the consistency of our quantum simulations. 
717: We stress that a good agreement between the numerical data 
718: of Fig.\ref{fig:h08exmpla} and the theoretical  spectral function (\ref{eq:Fw_renorm}) 
719: takes place in the \emph{whole} frequency range of $\omega _{m}$. 
720: This is, in fact, a very important consistency
721: check of the good relaxation of our paths ensemble at all frequencies,
722: including paths fluctuations at the lowest frequency available in
723: the system.
724: 
725: 
726: \begin{figure}[ht]
727: \includegraphics[ clip,
728:   width=115mm,
729:   height=80mm,
730:   keepaspectratio,
731:   angle=90,
732:   origin=c]{fig3.ps}
733: \vskip -10mm
734: \caption{\label{fig:h08exmplb} The phonon dispersion law $\nu (k)$: open circles show 
735: the data obtained from the fit as it is shown in Fig.\ref{fig:h08exmpla} for $l=1\div30$,
736: and the straight line shows the fit given by Eq.(\ref{eq:DispLaw}). The chain
737: parameters are the same as in Fig.\ref{fig:h08exmpla}, the wave vector $k=\pi l/L$.}
738: \end{figure}
739: 
740: By fitting the data for different phonon modes $l=1\div30$ one
741: can extract the dispersion relation for phonons  $\nu(k)$ 
742: where $k=\pi l/L$ (see Fig.\ref{fig:h08exmplb}). The majority of data
743: points follow the straight line given by the formula
744: \begin{equation}
745: \nu ^{2}(k)=\nu _{0}^{2}+c^{2}k^{2}
746: \label{eq:DispLaw}
747: \end{equation}
748: where $\nu _{0}$ is the phonon frequency gap, and $c$ is the velocity
749: of sound. The fit of numerical data gives: $\nu_0^2=3.204\pm 0.026$,
750: $c^2=15.0\pm 0.7$ for $\hbar=0.8$; $\nu_0^2=3.706\pm 0.019$,
751: $c^2=13.3\pm 0.5$ for $\hbar=0.2$. These quantum data should be compared
752: with the classical case where $\nu_0^2=3.697$, $c^2=11.5$ ($\hbar=0$).
753: We note that there is a difference between the frequency of the first
754: spatial harmonic $\nu_1$ and the frequency gap value $\nu_0$ obtained
755: from the dispersion law. However this difference is small and comparable
756: with the statistical errors. For small $\hbar$ the parameters of the 
757: dispersion law converge to their classical values. 
758: 
759: The described approach allows to obtain a complete information about 
760: low-energy phonon excitations in the whole quasiclassical region 
761: $\hbar \lesssim 1$.
762: 
763: 
764: \subsection{Transition region $1\lesssim \hbar \lesssim 2$.}
765: \label{sub:Transition-region}
766: 
767: As it is seen from Fig.\ref{fig:overview}, this region corresponds
768: to the transition between two regimes $\hbar \lesssim 1$ and 
769: $\hbar \gtrsim 2$ where the dependence of the quantum excitations on the particle
770: location in the chain looks quite regular.
771: 
772: As it will be shown later the irregular behavior in the region
773: $1\lesssim \hbar \lesssim 2$ is related to a significant increase
774: of the density of instantons. At high density the interaction between
775: instantons becomes important and results in onset of new phonon branch 
776: at $\hbar>2$.
777: In this section we discuss the properties of instantons and phonons 
778: and obtain estimates for their frequencies. 
779: 
780: Let us start with a discussion of tunneling effects. For $\hbar < 1$
781: the contribution of tunneling to the spectral function is exponentially
782: small being proportional to $\exp (-const/\hbar )$. However at  $\hbar >1$ 
783: the tunneling probability becomes large and it gives a significant contribution
784: to the spectral function. The transition to this regime is seen in  Fig.\ref{fig:overview} 
785: as a sequence of sharp isolated peaks. 
786: Following the pioneering paper \cite{Poly77}, a tunneling event can
787: be associated to an \emph{instanton}. In the imaginary time representation,
788: the instanton is a local jump between two wells, which is fast compared
789: to the mean time interval between subsequent jumps: while the size
790: of instantons (in time $\tau$) is practically independent of $\hbar $, the
791: separation between them is exponentially large in the quasiclassical
792: limit (low instanton density). In particular, this means that in the 
793: first approximation one may consider instantons as independent jumps, 
794: as can be also checked from a direct examination of our path ensemble.
795: Here we should stress two important properties of instantons
796: in the quantum FK chain:
797: 
798: (i) each jump of a particle $i$ in its 
799: position $x_i$ gives displacements of neighboring particles,
800: which decay exponentially with the distance from the jump
801: location, i.e. instantons are exponentially localized in space inside 
802: the chain;
803: 
804: (ii) instantons are distributed inhomogeneously along the chain since
805: the tunneling probability is highly sensitive to variations of barrier
806: heights due to chain inhomogeneity. 
807: 
808: A simple explanation of the exponential localization of an instanton
809: (along the chain) comes from the fact that the new static configuration
810: produced by it can be seen as a local \emph{static} defect on the
811: original configuration, which is known to dye away exponentially with
812: the classical Lyapunov exponent \cite{Au78,Au83b,Au83c}. Indeed,
813: in Fig.\ref{fig:overview} one can see that at $\hbar \sim 1.2\div 1.3$
814: instanton contributions are exponentially peaked around some particular
815: positions along the chain. 
816: 
817: Let us now consider the properties of elementary excitations originated
818: by instantons. There is a question how to select numerically a single instanton 
819: excitation.  Obviously, the ansatz (\ref{eq:profV}) used for phonons is good 
820: only for extended modes while instantons are localized in space. Therefore 
821: in this case we analyze numerically the frequency spectrum of a given
822: particle $i$ in the chain (defined by Eq.(\ref{eq:four_x})). If
823: $\hbar $ is not too high (close to one), instantons do not overlap, and the
824: main contribution to the spectrum comes from the instanton which is near to the given 
825: particle. This contribution reaches its maximum for a particle which actually
826: jumps.
827: 
828: At the same time besides instanton jumps the quantum motion of a particle 
829: in the chain contains a contribution of many phonons with different frequencies 
830: (see, e.g. Eq.(\ref{eq:Fw_x_i})). This phonon background should be subtracted in 
831: order to single out the contribution of instanton.  
832: Fortunately the frequencies $\nu _{l}$ of phonon excitations are much higher than 
833: the frequency $\nu ^{(inst)}$ of chosen instanton.
834: Hence these two types of excitations are well separated in the frequency domain.
835: Therefore in our analysis of the instanton contribution to the spectral function 
836: $F_i(\omega_m)\equiv\langle \left| a_{i}(\omega _{m})\right|^2\rangle$ 
837: we can restrict ourselves to the frequency domain $\omega _{m}<\omega _{bound}$,
838: where the boundary $\omega _{bound}$ is chosen by the condition 
839: $\nu ^{(inst)}\ll \omega _{bound}\ll \nu _{l}$
840: for all phonon modes $l$. In this frequency domain the phonon contribution
841: (\ref{eq:Fw}) is practically independent of $\omega _{m}$ and
842: can be replaced by some constant $C_{i}$. Thus we can extract
843: the frequency $\nu ^{(inst)}$ and the weight $f_{i}^{(inst)}$ of
844: the instanton excitation by following fit for the spectral function 
845: $F_{i}(\omega _{m})$:
846: 
847: \begin{equation}
848:  F_{i}(\omega _{m})=f_{i}^{(inst)}\hbar /\left(\omega _{m}^{2}+
849:  (\nu _{i}^{(inst)})^{2}\right)+C_{i}.\label{eq:fitInst}
850: \end{equation}
851: This fit contains three free parameters $\nu _{i}^{(inst)}$, $f_{i}^{(inst)}$ and
852: $C_{i}$. 
853: 
854: If instantons do not overlap then one may expect that
855: there are groups of particles which motion is dominated by a single
856: instanton. Inside each group the fit (\ref{eq:fitInst})
857: should give the same values for the frequencies $\nu _{i}^{(inst)}$, 
858: while the variation of weight $f_{i}^{(inst)}$ with $i$ 
859: determines the instanton profile along the
860: chain. This case is illustrated in Fig.\ref{fig:xEinst}a, which
861: corresponds to the early onset of instanton contribution at $\hbar =1.2$.
862: The six peaks in the bottom part of Fig.\ref{fig:xEinst}a show six non 
863: overlapping instantons, while the top part shows the corresponding frequencies 
864: as a function of particle index $i$ inside the chain. The peaks have different 
865: amplitudes, and the highest three of them involve groups of three particles
866: which have the same frequency inside each group. 
867: 
868: At higher $\hbar $
869: the number of instantons starts to grow rapidly and they begin to overlap.
870: A direct confirmation of this trend is seen in Fig.\ref{fig:xEinst}b
871: which corresponds to $\hbar =1.8$. Here all instantons have about ten percent
872: overlap with their neighbors and their interaction is rather
873: strong. As a result the step-like structure of frequencies, seen at
874: the top of the Fig.\ref{fig:xEinst}a, is practically destroyed. Thus
875: the instantons are ``collectivized'' and the consideration
876: of a single instanton as one-particle jump over barrier becomes 
877: not adequate. At higher $\hbar$ values this process leads to appearance
878: of a new phonon mode.  
879: %
880: \begin{figure}[ht]
881: \includegraphics[  width=115mm,
882:   height=85mm,
883:   keepaspectratio,
884:   angle=90,
885:   origin=c]{fig4.ps}
886: \vskip-12mm
887: \caption{\label{fig:xEinst}Dependence of the instanton frequency 
888: $\nu _{i}^{(inst)}$ (top)
889: and its weight $f_{i}^{(inst)}$ (bottom) on the position $i$ inside the chain 
890: (see text for explanations). (a) $\hbar=1.2$. Instantons do not overlap: 
891: the three highest peaks involves a group of three particles each. 
892: (b) $\hbar=1.8$ Instantons overlap. Chain parameters are
893: $s/r=34/21$, $K=5$, $\tau _{0}=320$.}
894: \end{figure}
895: 
896: 
897: %
898: \begin{figure}[th]
899: \includegraphics[  width=115mm,
900:   height=85mm, 
901:   keepaspectratio,
902:   angle=90,
903:   origin=c]{fig5.ps}
904: \vskip -12mm
905: \caption{\label{fig:spIP}Rescaled spectral function 
906: $F^{R}(\omega ,k_{1})=(\nu _{1}^{2}/\hbar )F(\omega ,k_{1})$
907: at the very beginning of instanton onset. The bump which appears at 
908: $\omega\lesssim 0.2$ corresponds to instanton admixture to the phonon 
909: spectral function. Black points, open circles and stars
910: correspond to $\hbar=1,1.1$ and $1.2$.
911: Lines show the fit to numerical data with Eq.(\ref{eq:fitPhIn}). Chain parameters
912: used in simulations are: $s/r=34/21$, $K=5$,$\tau _{0}=320$.}
913: \end{figure}
914: 
915: 
916: Let us now discuss the phonon properties in the region
917: $1\lesssim \hbar \lesssim 2$. As in the previous Section \ref{sub:Quasiclassical},
918: we extract them from the the spectral
919: function $F(\omega _{m},k_l)\equiv F^{(l)}(\omega _{m})$ obtained on the basis 
920: of the ansatz (\ref{eq:profV}) for a phonon mode $l$.  
921: However, in this region of $\hbar $
922: the spectral function $F^{(l)}(\omega _{m})$ has an admixture of instantons,
923: which grows rapidly with $\hbar $: a change
924: of $\hbar $ from $1.1$ to $1.2$ results in more than ten times
925: of the admixture weight(see Fig.\ref{fig:spIP}). We note that 
926: in the absence of instanton contribution the rescaled 
927: phonon spectral function 
928: $F^{R}(\omega ,k_{l})\equiv(\nu _{l}^{2}/\hbar )F(\omega ,k_{l})$
929: plotted in Fig.\ref{fig:spIP} should have an
930: universal limit equal to unity independent of the value of $\hbar$.
931: Hence the increase of $F^{R}(\omega ,k_{l})$ at small $\omega$
932: stresses the important contribution of instantons.
933: These instantons have different frequencies and their contribution
934: to the spectral function can be rather complicated.  
935: However, we can use again the strong frequency separation between instanon 
936: and phonon excitations. Indeed, for $\omega _{m}\gtrsim \nu _{l}\gg \nu ^{(inst)}$, 
937: all instanton contributions have an universal behavior  
938: $\propto \omega _{m}^{-2}$. Therefore we may replace them by a single 
939: ``instanton contribution'' with some average instanton frequency $\bar{\nu}^{inst}$. 
940: Then the spectral function can be fitted by a sum of two contributions:
941: \begin{eqnarray}
942:   F(\omega_{m},k_{l}) & = & f_{ph}\hbar /\left(\omega_{m}^{2}+
943:   \nu_l^{2}(k_{l})\right)\nonumber \\
944:    &  & +f_{inst}\hbar /\left(\omega_{m}^{2}+(\bar{\nu}^{inst}(k_{l})^2)\right)
945: \label{eq:fitPhIn}
946: \end{eqnarray}
947: where $f_{ph}(k),\nu_l(k)$ and $f_{inst}(k),\bar{\nu}^{in}(k)$ are
948: free fit parameters for phonons and instantons, respectively. 
949: 
950: Fitting the data for different phonon modes at $l=1-30$ we extract
951: the phonon dispersion law (\ref{eq:DispLaw}) (see Fig.\ref{fig:DLinTR},
952: compare with fit procedure for $\hbar<1$).
953: Contrary to the case of Fig.\ref{fig:h08exmplb} at $\hbar<1$, now
954: the data for the dispersion law $\nu (k)$ are scattered inside  
955: some finite band.
956: This indicates that the anzatz (\ref{eq:profV})
957: for the phonon profile is not so good to single out particular phonon
958: modes. Actually, the width of the
959: band provides some measure of the inaccuracy. Nevertheless, the phonon
960: modes are still approximately defined in the domain of $\hbar $ under consideration.
961: Their frequency decreases as $\hbar \to 2$ but remains separated
962: from zero by a finite gap.
963: 
964: %
965: \begin{figure}[th]
966: \includegraphics[  clip,
967:   width=110mm,
968:   height=85mm,
969:   keepaspectratio,
970:   angle=90,
971:   origin=c]{fig6.ps}
972: \vskip -10mm
973: \caption{\label{fig:DLinTR} Frequency $\nu (k)$ of phonons
974: versus wavenumber $k$ obtained from the fit (\ref{eq:fitPhIn})
975: at $\hbar =1.5$ (in the middle of the transition region).
976: Chain parameters used in simulations
977: are: $s/r=89/55$, $K=5$,$\tau _{0}=80$.}
978: \end{figure}
979: 
980: We also note that the fit (\ref{eq:fitPhIn}) allows formally
981: to determine the dispersion law $\nu _{inst}(k)$ for the instanton branch.
982: However the numerical data give irregular scattering of points inside a band
983: $0 \leq \nu_{inst}^2 \lesssim 0.01$ without any clear dependence on $k$. 
984: %(see Fig.\ref{fig:DLinTR}b).
985: The reason of such behavior is simple: projections of irregular positions of instantons
986: on the harmonic ansatz (\ref{eq:profV}) produce random weights for
987: contributions of different instantons. This result represents another 
988: manifestation of glass-like structure formed by instantons frozen/pinned
989: inside the chain. Since the positions of instantons are random the phonons
990: cannot propagate along the chain on large distances. In fact they become
991: localized by disorder in a way similar to the one-dimensional Anderson localization
992: (more details on the phonon properties in this regime will be presented elsewhere
993: \cite{PhonFK}).  
994: 
995: 
996: 
997: \subsection{New sliding phonon branch at $\hbar >2$}
998: \label{sub:New-phon}
999: 
1000: \begin{figure}[ht]
1001: \includegraphics[  width=115mm,
1002:   height=85mm,
1003:   keepaspectratio,
1004:   angle=90,
1005:   origin=c]{fig7.ps}
1006: \caption{\label{fig:path23} A sample of quantum paths of particles inside 
1007: some chain fragment, which corresponds to periods of the external potential 
1008: with numbers $22\div30$. Dashed lines show bottoms of the wells, thick solid
1009: lines show the tops of the barriers.
1010: Note an example of highly correlated instanton transitions at $\tau=15-30$
1011: which involve particles  in up to 5 periods of the potential.
1012: Parameters of the simulation are: $\hbar=2.3$, $s/r=89/55$, $K=5$ and 
1013: $\tau _{0}=80$. }
1014: \end{figure}
1015: 
1016: From Fig.\ref{fig:overview} one can see that the variation 
1017: of the amplitude $a_{i}(\omega_1)$ with $i$ (low frequency excitations)
1018: becomes rather smooth at $\hbar> 2$. 
1019: This means that the instantons are strongly overlapped here.
1020: In fact, this means 
1021: that several particles of some chain fragment jumps from one well to the 
1022: next simultaneously (see Fig.\ref{fig:path23}). 
1023: This is nothing but sliding of a local chain fragment
1024: along the periodical potential. The typical size of such fragments should
1025: grow with $\hbar $. If their sizes reach the size of the chain  then the sliding
1026: mode becomes open, and the phonon gap disappears. At this point the pinned
1027: instanton glass turns into the sliding phonon gas.
1028: 
1029: %
1030: \begin{figure}[th]
1031: \includegraphics[  width=115mm,
1032:   height=85mm,
1033:   keepaspectratio,
1034:   angle=90,
1035:   origin=c]{fig8.ps}
1036: \vskip -10mm
1037: \caption{\label{fig:DLinSR} The dispersion law $\nu (k)$ for sliding 
1038: phonons at $\hbar =2.5$. The numerical data (circles with error bars)
1039: are obtained from the fit (\ref{eq:fitPhIn}) for the instanton branch
1040: $\nu_{inst}(k_l)$.
1041: The straight line shows the best fit (\ref{eq:DispLaw}) to numerical data. The chain
1042: parameters used in simulations are: $s/r=89/55$, $K=5$,$\tau _{0}=80$.}
1043: \end{figure}
1044: 
1045: 
1046: A confirmation of this picture is presented in Fig.\ref{fig:DLinSR},
1047: which corresponds to $\hbar =2.5$ being just above the transition point
1048: $\hbar_c\approx 2$. The numerical data for the dispersion law $\nu (k)$
1049: in Fig.\ref{fig:DLinSR} are obtained from the fit (\ref{eq:fitPhIn}) 
1050: at $\hbar >2$. Here the behavior of phonon and instanton modes changes 
1051: dramatically: phonons data $\nu (k)$ are now irregularly scattered over a wide 
1052: band while the data points for instanton branch follow a single line, 
1053: reproducing fairly well a \emph{phonon}-like dispersion law with  zero gap
1054: (see Figs.\ref{fig:DLinSR},\ref{fig:Esurf}).
1055: In particular, the fit (\ref{eq:DispLaw}) gives 
1056: $\nu _{0}=0.04\pm 0.01$, which is close to zero. In fact, 
1057: this value is smaller 
1058: than the minimal frequency $2\pi /\tau _{0}$ ($\approx 0.079$, 
1059: at $\tau _{0}=80$) and therefore it is compatible with zero. 
1060: 
1061: On the contrary, the wide scattering of data points for the phonon
1062: branch indicates that the ansatz (\ref{eq:profV}) is not good for phonon 
1063: contribution at $\hbar\gtrsim 2$ (see Fig.\ref{fig:Esurf}). This scattering of points
1064: is related to the localization of high-frequency
1065: phonon modes. In contrast, a smooth behavior of data points for the instanton
1066: branch demonstrates that the instanton wave functions become close
1067: to the harmonic wave ansatz (\ref{eq:profV}). Hence these excitations
1068: are delocalized. This leads us to the conclusion, that 
1069: for $\hbar>2$ the instanton branch is replaced by a new gapless
1070: branch of new sliding phonons. 
1071: 
1072: 
1073: \subsection{Global picture of elementary excitations in the FK chain}
1074: \label{sub:SummaryExcit}
1075: 
1076: The ensemble of data for the dispersion law $\nu(k)$ of elementary excitations
1077: at different values of $\hbar$ is shown in Fig.\ref{fig:Esurf} by the two sheets
1078: representing the phonon branch (1) and the instanton branch (2). 
1079: The numerical data are obtained with the ansatz (\ref{eq:profV}) by the fit 
1080: (\ref{eq:fitPhIn}). 
1081: %
1082: \begin{figure}[ht]
1083: \includegraphics[clip,
1084:   width=120mm,
1085:   height=88mm,
1086:   keepaspectratio,
1087:   angle=270,
1088:   origin=c]{fig9.ps}
1089: \vskip -15mm
1090: \caption{\label{fig:Esurf} The frequency $\nu (k)$ of elementary excitations
1091: versus the wavenumber $k$ at different $\hbar $. Chain parameters used
1092: in simulations are: $s/r=89/55$, $K=5$,$\tau _{0}=80$. The two sheets (1) and (2)
1093: refer to phonon and instanton excitations respectively.} 
1094: \end{figure}
1095: 
1096: 
1097: The sheet (1) refers to phonons originated from classical phonon
1098: modes which are well reproduced in the limit $\hbar \to 0$. The frequencies
1099: of these modes are well separated from zero by a large gap.  Therefore
1100: one may say that  they  form the optical phonon branch. At
1101: $\hbar \gtrsim 1$ these modes show a tendency to become softer, and at 
1102: $\hbar \gtrsim 1.5$ their dependence on $k$ and $\hbar $ becomes irregular.
1103: As it was explained in the previous section this irregular dependence 
1104: is related to the glass-like structure of randomly pinned instantons
1105: which density increases with the growth of $\hbar$.
1106: The sheet (2) appears at $\hbar <2$ from the instanton contributions
1107: into the Feynman path integral. For $\hbar>\hbar_c\approx 2$
1108: this instanton branch turns into a new gapless branch of sliding phonons
1109: (see the discussion related to Fig.\ref{fig:DLinSR}).
1110: 
1111: %
1112: \begin{figure}[h]
1113: \includegraphics[  width=115mm,
1114:   height=85mm,
1115:   keepaspectratio,
1116:   angle=90,
1117:   origin=c]{fig10.ps}
1118: 
1119: \caption{\label{fig:PhSummary} Dependence of the phonon gap $\nu _{0}$ 
1120: and sound 
1121: velocity $c$ on $\hbar $ for the case of Fig.\ref{fig:Esurf} at $K=5$.
1122: Open symbols correspond to the sheet (1) in Fig.\ref{fig:Esurf}.
1123: Circles and triangles are obtained at $\tau_0=80$ for chain sizes 
1124: $s/r=89/55$ and $233/144$ respectively. Stars correspond
1125: to $s/r=34/21$ and $\tau _{0}=320$. The solid
1126: line (bottom) gives the fit (\ref{eq:CritExpC2}) for  $1.3 \leq\hbar\leq 2$
1127: (see text). The full circles (top) for $\hbar>2$ refer to the sliding 
1128: phonon branch from the sheet(2) of Fig.\ref{fig:Esurf}, they 
1129: indicate a zero phonon gap.}
1130: \end{figure}
1131: 
1132: 
1133: A more quantitative picture can be obtained from the
1134: numerical data for the gap $\nu _{0}$ and the sound velocity $c$ for 
1135: both sheets in Fig.\ref{fig:Esurf}.
1136: For the optical phonon branch the values of $\nu_0$ and $c$ are determined
1137: from the fit (\ref{eq:fitPhIn}) for different values of $\hbar$ 
1138: (see Fig.\ref{fig:PhSummary}).  
1139: The data show that the phonon gap remains finite and 
1140: large at $\hbar<\hbar_c\approx 2$. 
1141: In contrast, the sound velocity $c$ drops to zero as $\hbar$ 
1142: approaches the value $\hbar_c$. This decay is compatible with
1143: the fit 
1144: \begin{equation}
1145:   c^{2}=a(\hbar _{c}-\hbar )^{\alpha }
1146: \label{eq:CritExpC2}
1147: \end{equation}
1148: shown by the solid line with $a=20.5\pm 1.3$, $\hbar _{c}=2.0\pm 0.1$
1149: and the critical exponent $\alpha =0.52\pm 0.07$.
1150: Even if the numerical data for $c$ have certain fluctuations they still 
1151: clearly indicate the quantum phase transition at $\hbar_c\approx 2$,
1152: where the sound velocity $c$ drops to zero. Further extensive numerical studies 
1153: are required to determine the behavior in the vicinity of transition
1154: in a more precise way.
1155: %
1156: \begin{figure}[h]
1157: \includegraphics[  width=115mm,
1158:   height=85mm,
1159:   keepaspectratio,
1160:   angle=90,
1161:   origin=c]{fig11.ps}
1162: \caption{\label{fig:InSummary} Dependence of sound velocity $c$
1163: on $\hbar $ for excitations on the soft phonon sheet (2) 
1164: in Fig.\ref{fig:Esurf}. Points correspond to chain parameters $s/r=89/55$ 
1165: $K=5$ and $\tau _{0}=80$. The solid line shows the linear fit to the
1166: sound velocity data inside the region $\hbar >2$ (see text).}
1167: \end{figure}
1168: 
1169: 
1170: For the sliding phonon branch  we extract the parameters $\nu_0$ and $c$ from the data of 
1171: sheet (2) in Fig.\ref{fig:Esurf} using a more general fit 
1172: given by 
1173: \[
1174:   \nu ^{2}(k)=\nu _{0}^{2}+c^{2}k^{2}/(1+k^{2}/k_{B}^{2}).
1175: \]
1176: Compared to the standard case (\ref{eq:DispLaw}) 
1177:  we introduce an additional parameter $k_B$
1178: to take into account the saturation of $\nu(k)$ at large $k$ (see, 
1179: Fig.\ref{fig:DLinSR},\ref{fig:Esurf}).
1180: The numerical data show that the gap 
1181: $\nu _{0}$ is small and does not exceed the minimal frequency in 
1182: the system $\varpi =2\pi /\tau _{0}$. Hence, the gap $\nu_0$ is compatible with zero
1183: (see Fig.\ref{fig:PhSummary} (top)). 
1184: At the same time  
1185: the sound velocity $c$ for the sliding phonon branch grows approximately
1186: linearly with $\hbar $ (see Fig.\ref{fig:InSummary}). 
1187: The best fit of numerical data for $\hbar>2$ gives 
1188: \begin{equation}
1189:   c^{2}=a\cdot (\hbar -b),\label{eq:C2ofHbar}
1190: \end{equation}
1191: with $a=7.6\pm 0.1$, $b=1.57\pm 0.05$. Formally, the value of $b$ is different
1192: from the value $\hbar_c=2.0$ in (\ref{eq:CritExpC2}). However in view of large statistical
1193: fluctuations both fits for optical phonons and sliding phonons are compatible with the 
1194: quantum phase transition at $\hbar_c\approx 2$.
1195: 
1196: \subsection{Time correlations}
1197: \label{sub:Time-corr}
1198: 
1199: Time correlations are closely related to the frequency spectrum of
1200: elementary excitations in the system. Their analysis is probably the
1201: most traditional way to extract properties of the elementary excitations,
1202: see, e.g. \cite{Creu81,Shur84}. Below we discuss the connection
1203: of this traditional method with 
1204: our approach based on the Fourier spectrum of Feynman paths
1205: described above. We compare the results obtained by these two different methods.
1206: 
1207: The one-particle time correlators are directly related to the Fourier spectrum
1208: of Feynman paths, and are given by the following expression:
1209: \[
1210:  \left\langle x_{i}(0)x_{i}(\tau )\right\rangle =
1211:  \frac{1}{\sqrt{\tau _{0}}}\sum _{\omega _{m}}
1212:  \left|a_{i}(\omega _{m})\right|^{2}\exp \left(-i\omega _{m}\tau \right)
1213: \]
1214: In fact, many elementary excitations with different frequencies contribute 
1215: to a particle motion in the chain. In order to
1216: single out a particular phonon mode we study the correlators of
1217: normal modes $X^{l}(\tau )=\sum _{i}V_{i}^{(l)}x_{i}(\tau )$
1218: where $V_i^{(l)}$ are eigenvectors defined in Eqs.(\ref{eq:Om2})-(\ref{eq:Om3}).
1219: Then from Eq.(\ref{eq:corrAA}) we obtain
1220: \begin{eqnarray}
1221:  \left\langle X^{l}(0)X^{l}(\tau )\right\rangle  &  & =
1222:  \frac{1}{\sqrt{\tau _{0}}}\sum _{\omega _{m}}\left|A^{l}(\omega _{m})\right|^{2}
1223:  \exp \left(-i\omega _{m}\tau \right)\nonumber \\
1224:  = &  & \left\langle \left(X^{l}(0)\right)^{2}\right\rangle 
1225:  \left(\textrm{e}^{-\nu _{l}\tau }+
1226:  \textrm{e}^{-\nu _{l}(\tau _{0}-\tau)}\right),
1227: \label{eq:CorrPhT}
1228: \end{eqnarray}
1229: where $\left\langle \left(X^{l}(0)\right)^{2}\right\rangle =\hbar /2\nu _{l}$
1230: is the contribution of a single phonon mode and the periodicity along
1231: the time torus results in a second exponential term in (\ref{eq:CorrPhT}). 
1232: 
1233: However, Eq.(\ref{eq:corrAA}) assumes only small 
1234: quantum fluctuations ($\sim \hbar $) around some \emph{classical} trajectory. 
1235: But due to tunneling effects (or instantons) a particle jumps from one
1236: \emph{classical} trajectory to another and its actual motion
1237: is given by a sum of phonon $x_{i}^{(ph)}(\tau )$
1238: and instanton $x_{i}^{(inst)}(\tau )$ contributions:
1239: \[
1240:   x_{i}(\tau )=x_{i}^{(ph)}(\tau )+x_{i}^{(inst)}(\tau ).
1241: \]
1242: In general, both motions influence each other, but in the quasiclassical
1243: limit $\hbar \to 0$ they can be considered as independent. In this limit they
1244: have quite different frequency scales: the phonon
1245: frequency $\nu_l$ is of the order of $K^{1/2}$ while the frequency of tunneling jumps
1246: $\nu_{inst}$
1247: is exponentially small. In contrast, while the amplitude of phonon
1248: oscillations in the limit $\hbar \to 0$ is small: 
1249: $\left\langle \left(x_{i}^{(ph)}(\tau )\right)^{2}\right\rangle \propto \hbar $,
1250: the amplitude of jumps is defined by difference between equilibrium
1251: particle positions in two neighbor wells (in our case it is $\sim 3\div 4$),
1252: i.e. it does not depend on $\hbar $. Hence, for a jumping particle
1253: $\left\langle \left(x_{i}^{(inst)}(\tau )\right)^{2}\right\rangle \sim 10$
1254: which is not small even in the quasiclassical limit $\hbar \to 0$. 
1255: Therefore the instanton contribution to the time correlator 
1256: has the form \cite{Shur84}
1257: %\[
1258: % \left\langle x_{i}(0)x_{i}(\tau )\right\rangle _{inst}=
1259: % \left\langle \left(x_{i}^{(inst)}(\tau )\right)^{2}\right\rangle \times \\
1260: % \left(\textrm{e}^{-\nu _{inst}\tau }+ 
1261: % \textrm{e}^{-\nu _{inst}(\tau _{0}-\tau )}\right).
1262: %\]
1263: \begin{eqnarray}
1264:  \left\langle x_{i}(0)x_{i}(\tau )\right\rangle _{inst}=
1265:  \left\langle \left(x_{i}^{(inst)}(\tau )\right)^{2}\right\rangle \times \\
1266:  \left(\textrm{e}^{-\nu _{inst}\tau }+ 
1267:  \textrm{e}^{-\nu _{inst}(\tau _{0}-\tau )}\right). \nonumber
1268: \end{eqnarray}
1269: The pre-exponent factor is large for jumping particles even in the deep quasiclassical 
1270: regime.
1271: 
1272: In fact, not any particle can easily jump from one well to another: different
1273: classical trajectories have different actions and all jumps which
1274: result in a large change of action $\Delta S\gtrsim \hbar $ are
1275: inhibited. In particular, it is clearly seen in Fig.\ref{fig:overview}
1276: that the number of instanton peaks is smaller at smaller $\hbar$ since
1277: the contribution of transitions with large difference in action $\Delta S$
1278: is suppressed.
1279: 
1280: Finally, the general form for the time correlator 
1281: which takes into account the instantons contribution takes the form:
1282: \begin{eqnarray}
1283:  C(\tau ) & \equiv  & \left\langle X^{l}(0)X^{l}(\tau )\right\rangle =\nonumber \\
1284:  & = & \left\langle \left(X^{l}(0)\right)^{2}\right\rangle 
1285:  \left(\textrm{e}^{-\nu _{l}\tau }+\textrm{e}^{-\nu _{l}(\tau _{0}-\tau )}\right)
1286: \nonumber\\
1287:  & + & \sum _{inst}w_{in}\left(\textrm{e}^{-\nu _{inst}\tau }+
1288:  \textrm{e}^{-\nu _{inst}(\tau _{0}-\tau )}\right).
1289: \label{eq:CorrPhInT}
1290: \end{eqnarray}
1291: Here the first term describes the phonon contribution and in the second term
1292:  the sum is taken over instantons and the instanton weights 
1293: $w_{inst}=\sum _{i,k}V_{i}^{(l)}V_{k}^{(l)}
1294:  \left\langle x_{i}(0)x_{k}(0)\right\rangle $
1295: describe the overlap with the ansatz (\ref{eq:profV}). Both types 
1296: of contributions (\ref{eq:CorrPhInT})
1297: are clearly seen in Fig.\ref{fig:CorrTa} at $\hbar =1.4$. 
1298: The initial rapid drop at $\tau\lesssim 1.5$ 
1299: corresponds to phonon contribution while the slow decay
1300: at $\tau>1.5$ corresponds to the instanton contribution. 
1301: This initial drop is related to the existence of large
1302: quasiclassical gap for phonon excitations.
1303: For $\hbar=2.5>\hbar_c$ the gap disappears and the correlator decay
1304: very slowly (see Fig.\ref{fig:CorrTa}).
1305: \begin{figure}[h]
1306: \includegraphics[clip,
1307:   width=115mm,
1308:   height=85mm,
1309:   keepaspectratio,
1310:   angle=90,
1311:   origin=c]{fig12.ps}
1312: \caption{\label{fig:CorrTa} The numerically computed time correlator 
1313: $C(\tau )=\left\langle X^l(0)X^l(\tau )\right\rangle$ for $l=1$ at
1314: different time separations $\tau $ for $\hbar =1.4$ (lower points) and 
1315: $\hbar=2.5$ (upper points). The parameters of the chain are $s/r=89/55$ 
1316: $K=5$ and $\tau _{0}=80$.
1317: }
1318: \end{figure}
1319: 
1320: For $\hbar>\hbar_c$ we have a new phase where
1321: instantons are replaced by sliding phonons.
1322: Therefore in this regime we fit the numerical data for $C(\tau)$
1323: by Eq.(\ref{eq:CorrPhInT}) with $w_{inst}=0$. The results for 
1324: different $l$ allow to obtain numerically the dispersion law
1325: $\nu(k)$ shown in Fig.\ref{fig:CorrTb} (open circles). However
1326: the accuracy of this data is not so good compared to data
1327: (full circles) obtained 
1328: from the analysis of Fourier spectrum of Feynman paths described in the 
1329: previous sections.
1330:  
1331: \begin{figure}[h]
1332: \includegraphics[clip,
1333:   width=115mm,
1334:   height=85mm,
1335:   keepaspectratio,
1336:   angle=90,
1337:   origin=c]{fig13.ps}
1338: \caption{\label{fig:CorrTb} Frequency of elementary excitations $\nu (k)$ 
1339: versus $k$ for sliding phonon branch at $\hbar =2.5$. Open circles show 
1340: data extracted from the  time correlator $C(\tau)$ for different $l$ 
1341: (see Fig.\ref{fig:CorrTa}),
1342: full circles present the results obtained from the fit (\ref{eq:fitPhIn})
1343: for the frequency spectrum of Feynman paths (see Fig.\ref{fig:DLinSR}). 
1344: The chain parameters are the same as in Fig.\ref{fig:CorrTa}.
1345: }
1346: \end{figure}
1347: 
1348: 
1349: 
1350: \section{quantum phase transition}
1351: \label{sec:QFT}
1352: 
1353: The structual rearrangement of the elementary excitations spectrum can
1354: be related to \emph{quantum phase transition} in the chain from a
1355: ``pinned'' to ``sliding'' phase. However, we would like to stress that
1356: in contrast to the classical picture \cite{Au83b,Au83c} in the
1357: quantum case the absence of energy gap for excitations is not 
1358: necessarily related to the opening of the sliding phase. Indeed,
1359: due to quantum tunneling through Peierls-Nabbarro barriers related to
1360: instantons there are excitations
1361: with energy which decreases exponentially with the increase of barrier heights. 
1362: Formally, this corresponds to the disappearance of excitation gap. 
1363: Therefore, to confirm firmly
1364: the appearance of sliding phase we need to consider spatial correlations
1365: of particle motion in the chain. The sliding phase appears when
1366: the spatial correlation 
1367: length becomes comparable with the length of the chain.
1368: 
1369: 
1370: \subsection{Spatial correlation length}\label{sec:CorrLen}
1371: 
1372: The analysis of spatial correlations (correlations between the motion
1373: of different particles in the chain) is the most evident way to observe
1374: the transition between pinned and sliding phases. In principle, the
1375: spatial
1376: correlation function can be explicitly computed if the spectrum
1377: and the wave functions of elementary excitations are known. In fact, we
1378: have a complete quantitative picture for \emph{phonon} modes, at least
1379: for the low-lying ones. To obtain numerically the value of the spatial
1380: correlation length $l_c$ we assume that the elementary excitation
1381: spectrum is given by the dispersion relation 
1382: $\nu ^{2}(k)=\nu _{0}^{2}+c^{2}k^{2}$ 
1383: and the corresponding phonon modes are given the ansatz 
1384: (\ref{eq:profV}). Then the same-time spatial correlator reads:
1385: \begin{eqnarray}
1386:  \left\langle \left(x_{i}-\left\langle x_{i}\right\rangle \right)\left(x_{j}-
1387:  \left\langle x_{j}\right\rangle \right)\right\rangle  & = & 
1388:  \sum _{l}V_{i}^{(l)}V_{k}^{(l)}
1389:  \left\langle \left(X^{l}(0)\right)^{2}\right\rangle \nonumber \\
1390:  =\hbar \sum _{k}\frac{\cos (k(i-j))}{L\nu (k)} & = & 
1391:   \frac{\hbar }{\pi c}K_{0}\left(\left|i-j\right|\right/l_c),
1392: \label{eq:CorrX}
1393: \end{eqnarray}
1394: where $K_{0}(x)$ is McDonald's function, with a known asymptotics:
1395: $K_{0}(x)\rightarrow _{x\rightarrow \infty }\sqrt{\frac{\pi }{2x}}\textrm{e}^{-x}$,
1396: and $l_c= c/\nu _{0}$. 
1397: 
1398: 
1399: The fit (\ref{eq:CorrX}) of the numerical data gives the value
1400: of $l_c$  for different values of $\hbar $ as it is
1401: shown in Fig.\ref{fig:CorrX}. It is seen that the length $l_c$ has
1402: a sharp increase at $\hbar_c \approx 2$, and for $\hbar>\hbar_c$
1403: it becomes comparable with the length of the chain.
1404: This indicates that we have the \emph{quantum phase transition}
1405: near $\hbar_c\approx 2$.
1406: Indeed at $\hbar <2$ the length $l_c$
1407: is practically independent of the chain length $L$ while at $\hbar >2$
1408: it starts to increase with $L$. This is confirmed by the data of
1409: Fig.\ref{fig:CorrX} where in spite of strong fluctuations for $\hbar>2$
1410: the length $l_c$ becomes comparable with the chain size $L$.
1411:  
1412: \begin{figure}[H]
1413: \includegraphics[  width=65mm,
1414:   height=85mm,
1415:   angle=90,
1416:   origin=c]{fig14.ps}
1417: \caption{\label{fig:CorrX}The dependence of the spatial correlation length 
1418: $l_{c}$ on $\hbar $. Crosses,
1419: full and open circles correspond to chains with $s/r=34/21$,
1420: $89/55$ and $233/144$, respectively; $K=5$, $\tau _{0}=80$. 
1421: The vertical dot-dashed line marks the quantum phase transition
1422: at $\hbar_c\approx 2$.
1423: The positions of the horizontal dotted lines are proportional to the chain length 
1424: $L=2\pi r$.
1425: }
1426: \end{figure}
1427: 
1428: \subsection{Longwave response}
1429: 
1430: The appearance of the new sliding phonon phase implies
1431: that the response of amplitude of the longwave modes should be large
1432: in this regime. To test this expectation we present the dependence
1433: of the amplitude $A^l(\omega=0)$ on $\hbar$ in Fig.\ref{fig:X0sk}.
1434: The numerical data demonstrate a sharp increase of $A^l(0)$
1435: near $\hbar_c\approx 2$. It is interesting to note that due to the existence
1436: of frequency gap for phonon excitations at $\hbar<\hbar_c$ the amplitude $A^l(0)$
1437: is not very sensitive to the variations of $l$. On the contrary, for 
1438: $\hbar>\hbar_c$ the gap disappears and $A^l(0)$ starts to depend on $l$ 
1439: (see Fig.\ref{fig:X0sk}a). In a similar way $A^l(0)$ is independent of
1440: the chain length $L$ for $\hbar<\hbar_c$ while at $\hbar>\hbar_c$
1441: it grows with $L$ (see Fig.\ref{fig:X0sk}b).
1442: The numerical data of Fig.\ref{fig:X0sk}b for $1.5\leq\hbar\leq 2$
1443: can be described by the fit
1444: \begin{equation}\label{eq:gammaA2}
1445:   \langle (A^{l}(0))^2\rangle\approx A (\hbar_c-\hbar)^{-\gamma} 
1446: \end{equation} 
1447: which gives $\gamma=5.06\pm 1.72$ and $\hbar_c=2.01\pm0.05$. 
1448: 
1449: %
1450: \begin{figure}[H]
1451: \includegraphics[  width=115mm,
1452:   height=85mm,
1453:   keepaspectratio,
1454:   angle=90,
1455:   origin=c]{fig15.ps}
1456: \caption{\label{fig:X0sk}The amplitude square of zero-frequency quantum fluctuations
1457: $(A^{l}(0))^2$ as a function of $\hbar$ for $K=5$ and $\tau _{0}=80$. 
1458: (a) Full circles, squares
1459: and open circles are for modes $l=1,3$ and $5$, respectively,
1460: for the chain with $s/r=89/55$. (b) Open circles, full circles
1461: and stars are for $s/r=233/144$, $89/55$ and $34/21$, respectively;
1462: $l=1$. The vertical dot-dashed lines mark $\hbar_c\approx 2$,
1463: the horizontal dotted lines give the average values for $\hbar>2$. 
1464: }
1465: \end{figure}
1466: 
1467: \subsection{Other characteristics}
1468: 
1469: 
1470: Another way to test the transition from the pinned instanton glass to the sliding phonon phase
1471: is to measure the sensitivity to small shifts of boundaries of the chain. With this aim we consider the shift of boundary particles $i=0$ and $i=s$ given by 
1472: \begin{equation}
1473:    x_0(\tau)=a_S \cos(2 \pi\tau/\tau_0), \;\; x_s(\tau)=x_0+L,
1474: \end{equation} 
1475: where the amplitude $a_S$ was fixed at $a_S=0.5$. In Fig.\ref{fig:slide} we present the
1476: dependence of the response function 
1477: $R(i)=\langle x_0 (x_i-\langle x_i\rangle\rangle/\langle x_0^2 \rangle$
1478: on the particle number $i$ inside the chain.
1479: It is seen that the response in the center of the chain 
1480: drops strongly when the parameter $\hbar$ changes from 
1481: $\hbar=2.2$  to $\hbar=1.8$. This means that the chain is locked for 
1482: $\hbar<\hbar_c\approx 2$ while for $\hbar>\hbar_c$ the chain
1483: slides following the displacements of the boundary
1484: particles.
1485: \begin{figure}[ht]
1486: \includegraphics[  width=115mm,
1487:   height=85mm,
1488:   keepaspectratio,
1489:   angle=90,
1490:   origin=c]{fig16.ps}
1491: \caption{\label{fig:slide} Dependence of the response function $R(i)$ on 
1492: the particle position $i$
1493: in the chain at $\hbar=1.8$ (full circles) and $\hbar=2.2$ (open circles).
1494: The horizontal dashed 
1495: lines show the interval of averaging for the minimal response value
1496: $R_{min}$ in Fig.\ref{fig:CorrH}. Parameters of the chain
1497: are $s/r=34/21$, $K=5$ and $\tau _{0}=80$. }
1498: \end{figure}
1499: 
1500: 
1501: \begin{figure}[ht]
1502: \includegraphics[  width=115mm,
1503:   height=85mm,
1504:   keepaspectratio,
1505:   angle=90,
1506:   origin=c]{fig17.ps}
1507: \caption{\label{fig:CorrH} Dependence of the minimal response $R_{min}$ on $\hbar$
1508: for the parameters of Fig.\ref{fig:slide}. 
1509: }
1510: \end{figure}
1511: 
1512: In order to get a more quantitative picture, we estimate the value of the response
1513: function $R(i)$
1514: at its minimum in the middle of the chain by taking its average value inside the central region at $i=12\div22$: 
1515: $R_{min}=\langle R(i) \rangle_{i}$ (this interval is shown in Fig.\ref{fig:slide} by
1516: horizontal dashed lines). The dependence of $R_{min}$ on the parameter $\hbar$ 
1517: is presented in Fig.\ref{fig:CorrH}. We see, that the correlator in the central region 
1518: of the chain deviates from zero at $\hbar>2$. 
1519: According to the numerical data the response $R_{min}$ is very small for 
1520: $\hbar<\hbar_c\approx 2$ while it becomes rather strong for $\hbar>\hbar_c$.
1521: This confirms the existence of the quantum  phase transition 
1522: from the pinned to sliding phase at $\hbar=\hbar_c\approx 2$.
1523: \begin{figure}[ht]
1524: \includegraphics[  width=115mm,
1525:   height=85mm,
1526:   keepaspectratio,
1527:   angle=90,
1528:   origin=c]{fig18.ps}
1529: \caption{\label{fig:DEvsH} Dependence of the total
1530:  chain energy per particle 
1531: on $\hbar$ for the chain parameters $K=5$, $s/r=89/55$ and $\tau_0=80$;
1532: $\epsilon_0=5.302$ is an energy point chosen from convenience/illustration 
1533: reasons.}
1534: \end{figure}
1535: 
1536: What is the order of this transition? In our simulations we have repeated a cycle
1537: changing slowly $\hbar$ from $\hbar=1$ 
1538: to $\hbar=4$ and back about hundred times, but no hysteresis was found within
1539: the statistical errors. Indeed,   
1540: the difference of the chain energy per particle at the upward and backward paths 
1541: did not exceed 
1542: $10^{-3}$. This difference should be compared with the change of the energy per particle 
1543: $\delta E\approx 2.4$ which takes place when $\hbar$ changes from $\hbar=1$ to $\hbar=4$.
1544: The absence of hysteresis excludes  the phase transition of the first order. We 
1545: also do not see any breaks in the dependence of the chain energy on $\hbar$. 
1546: In order to make more visible small deviations 
1547: from a linear law we plot in Fig.\ref{fig:DEvsH} the quantity 
1548: $(\epsilon-\epsilon_0)/\hbar$, where $\epsilon_0=5.302$. The numerical data
1549: show  that the slope of energy dependence changes near $\hbar = \hbar_c \approx 2$.
1550: This change of slope is located  approximately at the same value of $\hbar$
1551: where the divergence of the correlation length takes place (see  Fig.\ref{fig:CorrX}).
1552: These data indicate that we have a second order quantum phase transition which appears
1553: near $\hbar_c\approx 2$.
1554: However, more extensive numerical simulations are required to determine more 
1555: precisely the order of the transition. 
1556: 
1557: 
1558: 
1559: \section{Conclusions}
1560: \label{sec:Conclusions}
1561: 
1562: We have studied quantum tunneling phenomena in a particular model
1563: of glassy material, the Frenkel-Kontorova chain. This system has
1564: a lot of states, which are exponentially degenerate and (meta)stable
1565: in the (quasi)classical limit. In the quantum case there are tunneling
1566: transitions between these states that can be understood in  terms 
1567: of instanton dynamics. In the quasiclassical limit at small $\hbar$ 
1568: the instanton density is small and instantons are local and isolated.
1569: With the increase of $\hbar $ their density grows and they start to overlap.
1570: 
1571: At sufficiently large $\hbar $ the instantons are coupled and
1572: become  collectivized. As a result, the tunneling of particles 
1573: in some fragments of the chain proceeds in correlated way.
1574: The size of these correlations grows until it reaches the size of the system.
1575: This leads to appearance of extended excitations
1576: and opening of a new gapless sliding phonon branch.
1577: Our data show that a quantum phase transition takes place
1578: between the pinned and sliding phases. Absence of hysteresis effects,
1579: as well as continuous dependence of chain energy on 
1580: $\hbar$ exclude the first order phase transition, so we can classify
1581: this transition as a continuous quantum phase transition. We stress
1582: that the quantum phase transition from pinned to sliding phase
1583: takes place in the regime where the classical chain always remains
1584: in the pinned phase with the finite phonon gap.
1585: 
1586: The direct analysis of Fourier spectrum of Feynman paths ensemble allowed
1587: to obtain detailed information on the dispersion law of low-lying 
1588: excitations in both quantum phases. Nevertheless, some questions remain open for further
1589: investigations. For example, one can analyze in more detail the effects 
1590: of interactions between instantons at low density and study their
1591: propagation properties in this regime. Another interesting remark 
1592: concerns the behavior of the system in the vicinity of the transition
1593: point at $\hbar_c\approx 2$. Indeed, in this region the kinetic
1594: energy per particle is approximately 0.6 that is about ten times smaller
1595: than the height of the potential barrier at $K=5$. Therefore more insights
1596: are required to understand the underlying physics of this transition.
1597:  
1598: One can ask on how general are our results obtained in the frame 
1599: of the Frenkel-Kontorova
1600: model? In fact, we never used any specific features of this model
1601: related to its nontrivial number theory properties. The only essential
1602: point is the existence  of an exponential number of quasidegenerate 
1603: states which is common for glassy materials and other disorder systems.
1604: Therefore it is very interesting to study an analogous quantum
1605: phase transition in systems with disorder and interactions.
1606: 
1607: This work was supported in part by the EC RTN network contract HPRN--CT-2000-0156;
1608: OVZ thanks the groups in Como and Toulouse for their hospitality
1609: during the work on this problem.
1610: %\begin{acknowledgments}
1611: %\end{acknowledgments}
1612: 
1613: %\bibliographystyle{revtex}
1614: %%%\bibliography{fk}
1615: 
1616: \begin{thebibliography}{99}
1617: \bibitem{FK38}Y.I. Frenkel and T.K. Kontorova, Zh. Eksp. Teor. Fiz. \textbf{8}, 1340 (1938)
1618: [Phys. Z. Sowjetunion \textbf{13}, 1 (1938)]. 
1619: \bibitem{Naba67} F. Nabarro, {\it Theory of Crystal Dislocations} (Clarendon, Oxford, 1967). 
1620: \bibitem{VPaAT84}V.L. Pokrovsky and A.L. Talapov, {\it Theory of Incommensurate Crystals}, 
1621: Soviet Scientifical Reviews Supplement Series Physics, vol.1 (Harwood, London, 1984). 
1622: \bibitem{SY71} S.C. Ying, Phys. Rev. B \textbf{3}, 4160 (1971). 
1623: \bibitem{Piet81}L. Pietronero, W.R. Schneider and S. Str\"{ a}sler, Phys. Rev. B \textbf{24},
1624: 2187 (1981). 
1625: \bibitem{Au78}S. Aubry, in \textit{Solitons and Condensed Matter Physics}, Eds. A.R.Bishop 
1626: and T.Schneider, (Springer, N.Y., 1978). 
1627: \bibitem{Au83a} S. Aubry, J. Phys. (France), \textbf{44}, 147 (1983).
1628: \bibitem{Flor96}L.M. Floria and J.J. Mazo, Adv. Phys. \textbf{45}, 505 (1996). 
1629: \bibitem{Brau97} O.M. Braun, T. Dauxois, M.V. Paliy, and M. Peyrard, Phys. Rev. Lett. 
1630: \textbf{78}, 1295 (1997); Phys. Rev. E \textbf{55}, 3598 (1997). 
1631: \bibitem{Cons00} L. Consoli, H.J.F. Knops, and A. Fasolino, Phys. Rev. Lett.
1632: {\bf 85}, 302 (2000).
1633: \bibitem{Borg89} F. Borgonovi, I. Guarneri and D. Shepelyansky, Phys. Rev. Lett.
1634: \textbf{63}, 2010 (1989).        
1635: \bibitem{Borg90} F. Borgonovi, I. Guarneri and D. Shepelyansky, Z. Phys. B \textbf{79}, 133 (1990).
1636: \bibitem{Borg89d} F. Borgonovi, Ph.D. thesis, Universita degli Studi di Pavia,
1637: Italia (1989).
1638: \bibitem{Berm94} G.P. Berman, E.N. Bulgakov and D.K. Campbell, Phys. Rev. B 
1639: \textbf{49}, 8212 (1994).
1640: \bibitem{Berm97} G.P. Berman, E.N. Bulgakov and D.K. Campbell, Physica D 
1641: \textbf{107}, 161 (1997).
1642: \bibitem{Hu98} B. Hu, B. Li and W.M. Zhang, Phys. Rev. E \textbf{58}, R4068 (1998).
1643: \bibitem{Hu99} B. Hu and B. Li, Europhys. Lett. \textbf{46}, 655 (1999).
1644: \bibitem{Hu01} B. Hu, B. Li and H. Zhao, Europhys. Lett. \textbf{53}, 342 (2001).
1645: \bibitem{Hu01r} B. Hu and  B. Li, Physica A \textbf{288}, 81 (2000).
1646: \bibitem{Ho01} C.-L. Ho and C.-I Chou, Phys. Rev. E \textbf{63}, 016203 (2001).
1647: \bibitem{Chir79} B.V.Chirikov, Phys. Rep. \textbf{52}, 263 (1979).
1648: \bibitem{Perc79} I.C.Percival, in {\it Nonlinear Dynamics and the Beam-Beam Interaction},
1649:         Eds. M.Month and J.C.Herra, AIP Conf. Proc., AIP, N.Y. {\bf 57}, 302 (1979).
1650: \bibitem{Au83b}S. Aubry, Physica D \textbf{7}, 240 (1983).
1651: \bibitem{Au83c}S. Aubry and P.Y. Le Daeron, Physica D \textbf{8}, 381 (1983).
1652: \bibitem{Mac84} R.S.~MacKay, J.~D.~Meiss and I.~C.~Percival, Physica D \textbf{13}, 55 (1984).
1653: \bibitem{Sch84} R. Schilling, Phys. Rev. Lett. {\bf 53}, 2258 (1984).
1654: \bibitem{Sch85} P. Reichert and R. Schilling, Phys. Rev. B \textbf{32}, 5731 (1985).
1655: \bibitem{Zhir01a} O.V. Zhirov, G. Casati and D.L. Shepelyansky, Phys. Rev. E \textbf{65}, 026220 (2002).
1656: \bibitem{Shur98} T. Shaefer and E.V. Shuryak, Rev. Mod. Phys. \textbf{70}, 323 (1998).
1657: \bibitem{Be80} H.U. Beyeler, L. Pietronero and S. Str\"{a}ssler, Phys. Rev. B \textbf{22}, 2988 (1980).
1658: \bibitem{PhonFK} O.V. Zhirov, G. Casati and D.L. Shepelyansky, 
1659: {\it Phonon localization in the Frenkel-Kontorova chain}, in preparation (2002).
1660: \bibitem{Poly77} A. Polyakov, Nucl. Phys. B \textbf{120}, 429 (1977).
1661: \bibitem{Creu81} M. Creutz and B. Freedman. Ann. Phys. \textbf{132}, 427 (1981).
1662: \bibitem{Shur84} E.V. Shuryak and O.V. Zhirov, Nucl. Phys. B \textbf{242}, 393 (1984).
1663: \bibitem{cQFT97} S.L. Sondhi, S.M. Girvin, J.P. Carini and D. Shahar, Rev. Mod. Phys. \textbf{69},
1664: 315 (1997).
1665: \bibitem{Metr53} N. Metropolis, A. Rosenbluth, M. Rosenbluth, A. Teller and E. Teller,
1666: J. Chem. Phys. \textbf{21}, 1087 (1953).
1667: 
1668: \end{thebibliography}
1669: 
1670: \end{document}
1671: